Next Article in Journal
Pd-Based Nano-Catalysts Promote Biomass Lignin Conversion into Value-Added Chemicals
Previous Article in Journal
Microstructure-Based Multiscale Modeling of Deformation in MarBN Steel under Uniaxial Tension: Experiments and Finite Element Simulations
Previous Article in Special Issue
Research Progress of Ba(Zn1/3Nb2/3)O3 Microwave Dielectric Ceramics: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phase Structure, Bond Features, and Microwave Dielectric Characteristics of Ruddlesden–Popper Type Sr2TiO4 Ceramics

1
School of Materials and Energy, University of Electronic Science and Technology of China, Chengdu 611731, China
2
China Zhenhua Group Yunke Electronic Co., Ltd., Guiyang 550018, China
3
China Zhenhua Group Xinyun Electronic Co., Ltd., Guiyang 550018, China
4
Academy of Advanced Interdisciplinary Research, Xidian University, Xi’an 710071, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(14), 5195; https://doi.org/10.3390/ma16145195
Submission received: 13 June 2023 / Revised: 11 July 2023 / Accepted: 22 July 2023 / Published: 24 July 2023
(This article belongs to the Special Issue Advances in Dielectric Ceramics and Their Applications)

Abstract

:
This work studied the phase constitution, bond characteristics, and microwave dielectric performances of Sr2TiO4 ceramics. Based on XRD and Rietveld refinement analysis, pure tetragonal Ruddlesden–Popper type Sr2TiO4 ceramic is synthesized at 1425~1525 °C. Meanwhile, the microstructure is dense and without porosity, indicating its high sinterability and densification. Great microwave dielectric performances can be obtained, namely an εr value of 39.41, and a Q × f value of 93,120 GHz, when sintered at 1475 °C. Under ideal sintering conditions, the extrinsic factors are minimized and can be ignored. Thus, the intrinsic factors are considered crucial in determining microwave dielectric performances. Based on the P–V–L complex chemical bond theory calculation, the largest bond ionicity, and proportions to the bond susceptibility from Sr–O bonds suggest that Sr–O bonds mainly determine the dielectric polarizability. However, the Ti–O bonds show lattice energy about three times larger than Sr–O bonds, emphasizing that the structural stability of Sr2TiO4 ceramics is dominated by Ti–O bonds, and the Ti–O bonds are vital in determining the intrinsic dielectric loss. The thermal expansion coefficient value of the Sr2TiO4 structure is also mainly decided by Ti–O bonds.

1. Introduction

With the development of 5G technology, microwave components, including microwave circuits, dielectric antennas, dielectric resonators, and dielectric filters, are widely used for their advantages in terms of small size, lightweight nature, and large quality factor (Q) value [1,2]. The microwave components fabricated by microwave dielectric ceramic exhibit great potential and market application value. Theoretically, an ideal ceramic candidate should have an adjustable dielectric constant (εr), an ideal dielectric loss (tanδ, tanδ = 1/Q), a high Q × f (Q × f refers to the Q value under a certain resonant frequency f) value, and a tunable temperature coefficient of resonance frequency (τf) value [3]. Hence, it is urgent to search for material candidates with outstanding microwave dielectric performance [4].
Ceramic systems with medium εr (20~70) values can reduce the device size and decrease the dissipation of microwave energy; as such, they have been attracting the attention of many scholars over the decades. Table 1 lists some popular ceramic materials that have been broadly investigated [5,6,7,8,9,10,11].
From Table 1, the niobate-based material systems have a large, coordinated range of dielectric constants and excellent Q × f value. Recently, a tetragonal Ruddlesden–Popper structure containing Sr2LaAlTiO7 (εr = 26.5, Q × f = 110,850 GHz, τf = 2.95 ppm/°C) [12], (Sr1−xCax)SmAlO4 (εr = 18.7, Q × f = 125,000 GHz, τf = −9 ppm/°C) [13], and Srn+1TinO3n+1 (an εr value of 42, a Q × f value~145,200 GHz, and a τf value ca. of 130 ppm/°C for n = 1) [10] has also been reported with great microwave dielectric performances.
It is widely acknowledged that the microwave dielectric performances of a ceramic system are primarily controlled by the phase structure when the extrinsic dielectric loss is properly regulated [14]. Even though the Sr2TiO4 ceramic shows an ultra-high Q × f value, the present literature reports have not conducted in-depth research on the structure–performance correlation. Thus, its dielectric characteristics concerning crystal structure are still unclear. Therefore, we wonder which part of the crystal structure influences the dielectric performances at the microwave range. Is it possible to determine the influence from the Sr site and Ti site and give guidance for ionic modification? To test this, Sr2TiO4 microwave dielectric ceramic was synthesized in the present study, where its structural impacts on chemical bond traits and microwave dielectric performances were investigated comprehensively.

2. Materials and Methods

Raw fine powders of SrCO3 (Aladdin, 99.9%) and TiO2 (Aladdin, 99.9%) were proportionally blended in view of the chemical formula of Sr2TiO4 in a ball-mill tank with zirconia balls and deionized water for 5 h. After that, the mixture was dried and sieved using a 120-mesh screen. Then, it was calcined at 1200 °C for 4 h to produce the Sr2TiO4 phase. The calcined powder was secondary ball-milled for 5 h. After that, the combination was dried and added with a polyvinyl alcohol solution to form cylinders (diameter: 12 mm; thickness: 6 mm), and then the pellets were sintered at 1425~1525 °C for 4 h.
The phase structure is analyzed by a powder X-ray diffraction instrument (Philips X’Pert Pro MPD, PANalytical, Morvern, UK). The structural Rietveld refinement examination was followed by using the GSAS-EXPGUI package to obtain crystal structural parameters, including cell volume, axis length, chemical bond type, chemical bond length, and chemical bond angle [15,16]. Before performing the analysis, the refined order was strictly required following the background, peak shape parameter fitting, and thermal vibration factor. The microstructure of sintered specimen was detected using scanning electron microscopy (SEM, FEI Inspect F), where the grain size distribution was analyzed by Nanomeasurer 1.2 software. The bulk density was acquired by Archimedes’ method. The theoretical and relative density were calculated based on the following [17]:
ρ t h e o = n × A N × V c e l l
ρ r e = ρ b u l k ρ t h e o
where n, A, Vcell, and N are the number of molecules in a unit cell (the number of structural units equals 2 for the Sr2TiO4 structure), molecular weight (g/mol), the unit cell volume (cm3), and Avogadro number (mol−1), respectively. The microwave dielectric performances were examined using a Hakki–Coleman dielectric resonator method under the TE011 mode with a network analyzer (E5071C, Agilent Technologies Co., Ltd., Santa Clara, CA, USA). The τf value was measured at 25 °C and 85 °C using the following formula:
τ f = f 85 f 25 60 f 25 10 6 ( ppm / ° C )
where f25 and f85 are the resonant frequency of ceramics at 25 °C and 85 °C, respectively.

3. Results

3.1. Phase Structure Investigation of Sr2TiO4 Ceramic

The X-ray diffraction (XRD) patterns of Sr2TiO4 ceramics sintered at 1425~1525 °C are shown in Figure 1a, and the XRD profile of Sr2TiO4 ceramic sintered at 1475 °C after the whole pattern fitting is presented in Figure 1b. All the observed diffraction peaks are assigned and indexed with JCPDS card No. 39-1471. No other peaks assigned to the second phase are noticed, suggesting a pure tetragonal Ruddlesden–Popper structure formation. The estimation criteria of Rwp = 5.99, Rp = 4.66, and χ2 = 2.008 are suitable and in an acceptable range, indicating the Rietveld refinement analysis’s validity. The refined structural parameters are a = b = 3.8898 (4) Å, c = 12.6089 (3) Å, and a Vcell of 190.779 (8) Å3 with an I4/mmm(139) space group.
The graphical representation of the Sr2TiO4 phase structure is shown in Figure 2. The Sr2TiO4 crystallizes in a tetragonal Ruddlesden–Popper structure with a space group of I4/mmm(139). It is treated as the arrangement of oxygen polyhedrons in an arrangement order of Ti–Sr–Sr–Ti–Sr–Sr–Ti along the c-axis direction. There are nine oxygen anions and six oxygen anions around the Sr2+ and Ti4+ cations, constructing a close-fitting bound [SrO9] and [TiO6] polyhedrons, where the [SrO9] polyhedrons are interconnected by sharing edges, and the [TiO6] octahedrons are interconnected by a shared vertex. The [SrO9] connects with [TiO6] through the edges. Particularly, three types of O anions occur in [SrO9]: Sr–O1 × 4 bonds, Sr–O2(1) × 4 bonds, and a Sr–O2(2) × 1 bond, while two types of O anions occur in [TiO6]: Ti–O1 × 4 bonds and Ti–O2 × 2 bonds. Exact atomic fractional coordinates and anion and cation spacing are listed in Table 2.

3.2. Microstructure Investigation of Sr2TiO4 Ceramic

The SEM images of Sr2TiO4 ceramic sintered at 1425~1525 °C are presented in Figure 3. It is shown that when the sintering temperature is as low as 1425 °C, some observable micropores are accompanied by a small average grain size of about 1.89 μm, indicating that the Sr2TiO4 ceramic is not well-densified. As the temperature gradually increases from 1425 °C to 1475 °C, as shown in Figure 3a–c, the growth of grain is promoted and the amounts of micropores decline. Specifically, high densification can be achieved in Figure 3c, which suggests its high sinterability. It also shows a uniform distribution of grain size, where the largest, smallest, and mean grain sizes are about 5.69 μm, 1.19 μm, and 2.58 μm, respectively. The bulk density obtained via the Archimedes method is 4.8352 g/cm3, about 96.74% of the theoretical density. However, as the temperature further increases to 1500 °C and 1525 °C, as shown in Figure 3d,e, it is observed that the size of the grain is abnormally large, about two times larger than that in 1475 °C. This phenomenon may result from the secondary grain growth caused by high sintering temperature, which is unfavorable for the uniformity of grain size distribution.

3.3. Bond Traits and Microwave Dielectric Performances Investigation of Sr2TiO4 Ceramic

The Sr2TiO4 ceramics are sintered at 1425~1525 °C, and the developments of densifications and microwave dielectric performances are shown in Figure 4. Firstly, from Figure 4a, it is found that the Sr2TiO4 ceramics reach high densifications (>95%) at 1475 °C. The variations of εr value are dominated by extrinsic and intrinsic factors, such as densification, dielectric polarizability, and phase compositions [18]. For instance, in this study, the dielectric polarizability remains unchanged since there is no additional ionic dopant. Moreover, the phase structure is the same tetragonal Ruddlesden–Popper type. Thus, the εr value in Figure 4b presents a comparable trend with the relative density. Figure 4c shows that the Q × f value increases to 93,120 GHz at 1475 °C and declines afterward. It is widely acknowledged that the Q × f value is sensitive to grain growth, densification, and phase constitutions [19]. In our present study, since the phase structure remains unchanged, the densification and growth of grain is important for the development of the Q × f value. Combined with the SEM images and variation in average grain size, it is clearly observed that the grains are fully grown, and the porosity of the microstructure is reduced, which is beneficial for reducing the number of grain boundaries per unit volume, reducing external grain boundary losses, and then increasing the Q × f value. However, excessive sintering temperature causes abnormal grain growth, which disrupts the uniformity of grain size and is not conducive to the improvement of Q × f value [20]. The variations in τf value in Figure 4d are similar to the evolutionary trends of the εr and Q × f values, which are also caused by the mutual influences from the densification, phase structure, and dielectric polarizability.
Figure 4 shows that optimum microwave dielectric properties can be acquired when sintered at 1475 °C: εr = 39.41, Q × f = 93,120 GHz, τf = 110.54 ppm/°C. In the optimum state, the impacts of extrinsic loss can be ignored; thus, the intrinsic factor determining the microwave dielectric performances of Sr2TiO4 ceramics, namely the structure–property connection, should be unambiguously analyzed. The chemical bond theory founded by Philips [21], Van Vechten [22], and Levine [23] (hereafter shortened as P–V–L complex chemical bond theory) deliver a strategy for calculating the fundamental chemical bonds characteristics, such as bond ionicity (fi), bond covalency (fc), bond susceptibility (χ), lattice energy (U), and the thermal expansion coefficient (αL). The phase structure characteristics are able to be specifically classified into chemical bond properties within the structure using these bond traits [4,24].
Generally, a crystal can be treated as a combination of chemical bonds between ions. The molecular formula of a crystal can also be regarded as the summation of chemical bonds. A chemical bond is binary, and the binary compounds AmBn can express its chemical formula; thus, the complex crystal AaBbCcDd is disassembled into the summation of binary crystals based on the crystal structure configuration, molecular formula, and chemical bond types [25]. Followed this, the traits of multi-bonds can be solved by treating them as a single bond. Thus, the binary expressions of Sr2TiO4 are firstly recognized on the basis of its unique phase structure, coordinating atmosphere, and charge distribution of ions, as shown in Figure 5.
It shall be noted that the effective valence electron numbers (Z) of the O anions in Sr–O and Ti–O bonds are 4 and 12, respectively. Based on the resolution theory of binary crystal formulas for multi-crystalline crystals proposed by Zhang [25], the summation of binary crystals in Sr2TiO4 is described as follows:
S r 2 T i O 4 = S r 2 T i O 1 2 O 2 2 S r 2 T i O 4 = S r 8 / 9 O 1 4 / 3 + S r 8 / 9 O 2 ( 1 ) 4 / 3 + S r 2 / 9 O 2 ( 2 ) 1 / 3 + T i 2 / 3 O 1 2 / 3 + T i 1 / 3 O 2 1 / 3
The chemical bonds are not 100% ionic or covalent for an ionic crystal. Thus, it is crucial to distinguish the ionic part and covalent part of chemical bonds in Sr2TiO4 ceramic. The bond ionicity value of any bond ( f i μ ) is estimated using the following equations:
f i μ = ( C μ ) 2 ( E g μ ) 2
f c μ = 1 f i μ
f c μ = ( E h μ ) 2 ( E g μ ) 2
where Cμ, E h μ , and E g μ are the heteropolar part, homopolar part, and average energy gap of Sr2TiO4 ceramic, respectively. They are calculated based on the following equations:
E g μ = ( E h μ ) 2 + ( C μ ) 2
E h μ = 39.74 r 2.48
C μ = 14.4 × b μ × ( Z A μ n m Z B μ ) × exp ( k s r 0 μ ) / r 0 μ   ( n >   m )
where r 0 μ equals half of the length dμ, acquired from the Rietveld refinement method; Z A μ and Z B μ signify the number of effective valence electrons; exp ( k s r 0 μ ) is the calculated Thomas–Fermi screening index; bμ is the correction index related with the mean coordination number N c μ , which is obtained as follows:
N c μ = m m + n N c A μ + n m + n N c B μ
b μ = 0 . 089 ( N c μ ) 1.48
where m and n are gained from the binary bonding formula expression AmBn, and N c A μ , N c B μ are the coordination numbers of the A and B atoms (for Sr, Ti, and O, they are 9, 6, and 6, respectively). The exp ( k s r 0 μ ) is obtained as follows:
k s μ = ( 4 k F μ π a 0 ) 1 / 2
where a0 is the Bohr radius, and k F μ the Fermi wave vector. The k F μ is calculated as follows:
k F μ = ( 3 π 2 N e μ ) 1 / 3
where N e μ is the effective valence electron density, as obtained from the following:
N e μ = n v μ v b μ = ( Z A μ N c A μ + Z B μ N c B μ ) V b μ
V b μ = ( d μ ) 3 ( d v ) 3 N b μ
where n v μ , V b μ , and N b μ are the number of valence electrons, bond volume, and bond density of Sr2TiO4 ceramic, respectively. Therefore, the f i μ value of chemical bonds is calculated, and the comparisons are shown in Figure 6.
It is found that Sr–O bonds show larger f i μ values in comparison to Ti–O bonds, demonstrating that Sr–O bonds in [SrO9] polyhedrons may provide a larger contribution to the dielectric polarizations within Sr2TiO4 ceramic. Nevertheless, the theoretical calculation of dielectric polarizations shall be evaluated to comprehend chemical bond contributions better. The bond susceptibility (χμ) in P–V–L complex chemical bond theory reflects the dielectric polarizations of any bond, which is defined as follows [23]:
χ μ = ( Ω p ) 2 4 π ( E g μ ) 2
χ = F μ × χ μ
where ℏ and Ωp represent Planck’s constant and plasma frequency of bond, respectively; Fμ is the proportion of μ type of bonds in all the bonds. The calculated χμ and proportions χμ/χ are listed in Table 3. It is also found that the largest contributions to the dielectric polarization come from Sr–O2(1) and Sr–O1 bonds. Moreover, Sr–O bonds contribute about 62.13% to the dielectric polarization, greater than the Ti–O bonds (37.87%). This result confirms the conclusion of bond ionicity analysis, showing that the Sr–O bonds may be more important in determining the dielectric polarization.
The Q × f value is closely connected with the dielectric loss tanδ (Q = 1/tanδ, f is the resonant frequency), which is considered from the inherent and extrinsic loss. Extrinsic loss refers to the size of grain growth, densification, and phase constitutions, etc. [26]. Internally, non-harmonicity of lattice vibrations in a faultless crystal produces intrinsic loss [27]. Lattice energy (U) indicates the binding abilities between ions [28]. A phase structure shows high stability with a strong binding ability [29]. The U value is analyzed as follows:
U t o t a l = μ ( U b c μ + U b i μ )
U b c μ = 2100 × m ( Z + μ ) 1 . 64 ( d μ ) 0 . 75 f c μ
U b i μ = 1270 ( m + n ) Z + μ Z μ d μ ( 1 0.4 d μ ) f i μ
where Z + μ and Z μ are the valence states of the Sr4+, Ti4+, and O2−, dμ is the distance between cation and anion, and the m/n value is gained from the binary bonding formula. The calculated U value and the contributions of chemical bonds are shown in Figure 7. As we can see, the UTi–O is approximately three times greater than USr–O, which emphasizes that the Ti–O bonds play a dominating role in the lattice stability and Ti–O bonds are more crucial in regulating the intrinsic dielectric loss.
The temperature coefficient of resonance frequency (τf) reflects the temperature stability of the ceramic system in variable conditions. It shall be properly adjusted based on practical application. From the previous literature, the τf value is affected by the effects of dielectric polarization capacity and the coefficient of thermal expansion as follows [30]:
τ f = ( α L + τ ε 2 )
where αL represents the thermal expansion coefficient, and τε stands for the permittivity temperature coefficient. The τf value is inversely proportional to the τε and αL values. The τε is typically affected by dielectric polarizability and increases with the decline in the εr value [31]. Based on the P–V–L complex chemical bond theory, the τf value is inversely proportional to the αL value. The αL is created by the anharmonicity in the Sr2TiO4 phase structure [32], which is calculated by using the estimated lattice energy U, as follows:
α = F m n × α m n μ
α m n μ = 3.1685 + 0.8376 γ m n
γ m n = k Z A N C A μ U ( A m B n ) Δ A β m n
β m n = m ( m + n ) 2 n
where Fmn is the proportion of this type of chemical bond in the Sr2TiO4 phase structure. k and ΔA are the Boltzmann constant and correction index, respectively. The results are presented in Table 4. It is found that the αL value of the Sr2TiO4 structure is estimated at about ~16.45 ppm/°C, and the Ti–O bonds are more significant than Sr–O bonds.

4. Conclusions

This study mainly investigates the crystal structure, bond characteristics, and structure–property relationship of Sr2TiO4 ceramics. The Sr2TiO4 ceramic sintered at 1475 °C shows a compact microstructure and great microwave dielectric performances, namely an εr value of 39.41, a Q × f value of 93,120 GHz, and a τf value ca. of 110.54 ppm/°C. Under the optimum sintering temperature, the extrinsic factors can be excluded, and the structure’s configuration mainly decides the structure–property relationship. On the basis of the P–V–L complex chemical bond theory, it is shown that the Sr–O bonds have larger average bond ionicity (fi) values (about 62.2538%) than Ti–O bonds (about 56.6779%). Also, the bond susceptibility (χμ) value of Sr–O bonds indicates their contribution of about 62.13% to the dielectric polarization, which is also greater than the Ti–O bonds (37.87%). This result demonstrates that the Sr–O bonds are the main factors contributing to the dielectric polarizability. The largest lattice energy (U) value of Ti–O bonds, however, about three times larger than Sr–O bonds, clarifies their significance in structural stability. The Ti–O bonds are also crucial for developing the thermal expansion coefficient value of the Sr2TiO4 structure, which is important for the development of the temperature coefficient of resonance frequency.

Author Contributions

Conceptualization, Writing-original draft, data curation, J.Y.; Writing-review & editing, data curation, J.P.; Formal analysis, funding acquisition, X.L.; Investigation, methodology, L.A.; Visualization, validation, Q.X.; Formal analysis, software, X.W.; Resources, supervision, H.Y.; Project administration, methodology, X.T. All authors have read and agreed to the published version of the manuscript.

Funding

This study is funded by the open research fund of Songshan Lake Materials Laboratory (No. 2022SLABFN20); the 2023 Guizhou Science and Technology Support Plan (Guizhou Science and Technology Support [2023] General 420); the Qinchuangyuan Citing High-level Innovation and Entrepreneurship Talent Projects (No. QCYRCXM-2022-40); the Natural Science Basic Research Program of Shaanxi (No. 2022JQ-390); the Fundamental Research Funds for the Central Universities (No. XJS222209).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wu, F.-F.; Zhou, D.; Du, C.; Jin, B.-B.; Li, C.; Qi, Z.-M.; Sun, S.; Zhou, T.; Li, Q.; Zhang, X.-Q. Design of a Sub-6 GHz Dielectric Resonator Antenna with Novel Temperature-Stabilized (Sm1–xBix)NbO4 (x = 0–0.15) Microwave Dielectric Ceramics. ACS Appl. Mater. Interfaces 2022, 14, 7030–7038. [Google Scholar] [CrossRef] [PubMed]
  2. Yang, H.; Chai, L.; Liang, G.; Xing, M.; Fang, Z.; Zhang, X.; Qin, T.; Li, E. Structure, far-infrared spectroscopy, microwave dielectric properties, and improved low-temperature sintering characteristics of tri-rutile Mg0.5Ti0.5TaO4 ceramics. J. Adv. Ceram. 2023, 12, 296–308. [Google Scholar] [CrossRef]
  3. Sebastian, M.T.; Jantunen, H. Low loss dielectric materials for LTCC applications: A review. Int. Mater. Rev. 2008, 53, 57–90. [Google Scholar] [CrossRef]
  4. Yang, H.; Zhang, S.; Yang, H.; Li, E. Usage of P–V–L bond theory in studying the structural/property regulation of microwave dielectric ceramics: A review. Inorg. Chem. Front. 2020, 7, 4711–4753. [Google Scholar] [CrossRef]
  5. Zhang, P.; Zhao, Y.; Liu, J.; Song, Z.; Xiao, M.; Wang, X. Correlation of crystal structure and microwave dielectric properties of Nd1.02(Nb1-xTax)0.988O4 ceramic. Dalton Trans. 2015, 44, 5053–5057. [Google Scholar] [CrossRef]
  6. Zhang, J.; Luo, Y.; Yue, Z.; Li, L. High-Q and temperature-stable microwave dielectrics in layer cofired Zn1.01Nb2O6/TiO2/Zn1.01Nb2O6 ceramic architectures. J. Am. Ceram. Soc. 2019, 102, 342–350. [Google Scholar] [CrossRef] [Green Version]
  7. Luo, W.; Li, L.; Yu, S.; Sun, Z.; Zhang, B.; Xia, F. Structural, Raman spectroscopic and microwave dielectric studies on high-Q materials in Ge-doped ZnTiNb2O8 systems. J. Alloys Compd. 2018, 741, 969–974. [Google Scholar] [CrossRef]
  8. Yang, H.; Zhang, S.; Yang, H.; Zhang, X.; Li, E. Structural Evolution and Microwave Dielectric Properties of xZn0.5Ti0.5NbO4-(1–x)Zn0.15Nb0.3Ti0.55O2 Ceramics. Inorg. Chem. 2018, 57, 8264–8275. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Zhang, Y.; Xiang, M. Crystal structure and microwave dielectric characteristics of Zr-substituted CoTiNb2O8 ceramics. J. Eur. Ceram. Soc. 2016, 36, 1945–1951. [Google Scholar] [CrossRef]
  10. Liu, B.; Li, L.; Liu, X.Q.; Chen, X.M. Srn+1TinO3n+1 (n = 1, 2) microwave dielectric ceramics with medium dielectric constant and ultra-low dielectric loss. J. Am. Ceram. Soc. 2017, 100, 496–500. [Google Scholar] [CrossRef]
  11. Tseng, C.-F. Microwave dielectric properties of a new Cu0.5Ti0.5NbO4 ceramics. J. Eur. Ceram. Soc. 2015, 35, 383–387. [Google Scholar] [CrossRef]
  12. Liu, B.; Liu, X.Q.; Chen, X.M. Sr2LaAlTiO7: A New Ruddlesden-Popper Compound with Excellent Microwave Dielectric Properties. J. Mater. Chem. C 2016, 4, 1720–1726. [Google Scholar] [CrossRef]
  13. Mao, M.M.; Chen, X.M. Infrared Reflectivity Spectra and Microwave Dielectric Properties of (Sr1−xCax)SmAlO4 (0 ≤ x ≤ 1) Ceramics. Int. J. Appl. Ceram. Technol. 2011, 8, 1023–1030. [Google Scholar] [CrossRef]
  14. Tian, H.; Zhou, X.; Jiang, T.; Du, J.; Wu, H.; Lu, Y.; Zhou, Y.Y.; Yue, Z. Bond characteristics and microwave dielectric properties of (Mn1/3Sb2/3)4+ doped molybdate based low-temperature sintering ceramics. J. Alloys Compd. 2022, 906, 164333. [Google Scholar] [CrossRef]
  15. Toby, B. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef] [Green Version]
  16. Larson, A.C.; Von Dreele, R.B. General Structure Analysis System (GSAS); Los Alamos National Laboratory Report LAUR 86-748; Los Alamos National Laboratory: Los Alamos, NM, USA, 2004.
  17. Takada, T.; Wang, S.F.; Yoshikawa, S.; Jang, S.-J.; Newnham, R.E. Effect of Glass Additions on BaO–TiO2–WO3 Microwave Ceramics. J. Am. Ceram. Soc. 1994, 77, 1909–1916. [Google Scholar] [CrossRef]
  18. Wang, S.; Zhang, Y. Structure, bond characteristics and microwave dielectric properties of new A0.75Ti0.75Ta1.5O6 (A = Ni, Co, Zn and Mg) ceramics based on complex chemical bond theory. J. Eur. Ceram. Soc. 2020, 40, 1181–1185. [Google Scholar] [CrossRef]
  19. Wang, G.; Zhang, D.; Huang, X.; Rao, Y.; Yang, Y.; Gan, G.; Lai, Y.; Xu, F.; Li, J.; Liao, Y.; et al. Crystal structure and enhanced microwave dielectric properties of Ta5+ substituted Li3Mg2NbO6 ceramics. J. Am. Ceram. Soc. 2020, 103, 214–223. [Google Scholar] [CrossRef] [Green Version]
  20. Fang, Z.; Tang, B.; Yuan, Y.; Zhang, X.; Zhang, S. Structure and microwave dielectric properties of the Li2/3(1−x)Sn1/3(1−x)MgxO systems (x = 0-4/7). J. Am. Ceram. Soc. 2018, 101, 252–264. [Google Scholar] [CrossRef]
  21. Phillips, J.C. Ionicity of the Chemical Bond in Crystals. Rev. Modern Phys. 1970, 42, 317–356. [Google Scholar] [CrossRef]
  22. Van Vechten, J.A. Quantum Dielectric Theory of Electronegativity in Covalent Systems. I. Electronic Dielectric Constant. Phys. Rev. 1969, 182, 891–905. [Google Scholar] [CrossRef]
  23. Levine, B.F. Bond susceptibilities and ionicities in complex crystal structures. J. Chem. Phys. 1973, 59, 1463–1486. [Google Scholar] [CrossRef]
  24. Tian, H.; Zheng, J.; Liu, L.; Wu, H.; Kimura, H.; Lu, Y.; Yue, Z. Structure characteristics and microwave dielectric properties of Pr2(Zr1−xTix)3(MoO4)9 solid solution ceramic with a stable temperature coefficient. J. Mater. Sci. Technol. 2022, 116, 121–129. [Google Scholar] [CrossRef]
  25. Wu, Z.J.; Meng, Q.B.; Zhang, S.Y. Semiempirical study on the valences of Cu and bond covalency in Y1-xCaxBa2Cu3O6-y. Phys. Rev. B 1998, 58, 958–962. [Google Scholar] [CrossRef]
  26. Zhou, X.; Liu, L.; Sun, J.; Zhang, N.; Sun, H.; Wu, H.; Tao, W. Effects of (Mg1/3Sb2/3)4+ substitution on the structure and microwave dielectric properties of Ce2Zr3(MoO4)9 ceramics. J. Adv. Ceram. 2021, 10, 778–789. [Google Scholar] [CrossRef]
  27. Kim, E.S.; Chun, B.S.; Freer, R.; Cernik, R.J. Effects of packing fraction and bond valence on microwave dielectric properties of A2+B6+O4 (A2+: Ca, Pb, Ba; B6+: Mo, W) ceramics. J. Eur. Ceram. Soc. 2010, 30, 1731–1736. [Google Scholar] [CrossRef]
  28. Liu, D.; Zhang, S.; Wu, Z. Lattice energy estimation for inorganic ionic crystals. Inorg. Chem. 2003, 42, 2465–2469. [Google Scholar] [CrossRef]
  29. Xia, W.-S.; Li, L.-X.; Ning, P.-F.; Liao, Q.-W. Relationship Between Bond Ionicity, Lattice Energy, and Microwave Dielectric Properties of Zn(Ta1−xNbx)2O6 Ceramics. J. Am. Ceram. Soc. 2012, 95, 2587–2592. [Google Scholar] [CrossRef]
  30. Bosman, A.J.; Havinga, E.E. Temperature Dependence of Dielectric Constants of Cubic Ionic Compounds. Phys. Rev. 1963, 129, 1593–1600. [Google Scholar] [CrossRef]
  31. Ma, P.P.; Liu, X.Q.; Zhang, F.Q.; Xing, J.J.; Chen, X.M. Sr(Ga0.5Nb0.5)1−xTixO3 Low-Loss Microwave Dielectric Ceramics with Medium Dielectric Constant. J. Am. Ceram. Soc. 2015, 98, 2534–2540. [Google Scholar] [CrossRef]
  32. Zhang, S.; Li, H.; Zhou, S.; Pan, T. Estimation Thermal Expansion Coefficient from Lattice Energy for Inorganic Crystals. Jpn. J. Appl. Phys. 2006, 45, 8801–8804. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of Sr2TiO4 ceramics sintered at 1425~1525 °C; (b) XRD profile of Sr2TiO4 ceramic sintered at 1475 °C after whole pattern fitting.
Figure 1. (a) XRD patterns of Sr2TiO4 ceramics sintered at 1425~1525 °C; (b) XRD profile of Sr2TiO4 ceramic sintered at 1475 °C after whole pattern fitting.
Materials 16 05195 g001
Figure 2. Schematic representation of Sr2TiO4 crystal structure.
Figure 2. Schematic representation of Sr2TiO4 crystal structure.
Materials 16 05195 g002
Figure 3. SEM images of Sr2TiO4 ceramics sintered at (a) 1425 °C, (b) 1450 °C, (c) 1475 °C, (d) 1500 °C, and (e) 1525 °C; (f) the variation in average grain size.
Figure 3. SEM images of Sr2TiO4 ceramics sintered at (a) 1425 °C, (b) 1450 °C, (c) 1475 °C, (d) 1500 °C, and (e) 1525 °C; (f) the variation in average grain size.
Materials 16 05195 g003
Figure 4. Development of (a) relative density, (b) εr value (measured in the ranges of 4.8~5.2 GHz), (c) Q × f value, and (d) τf value.
Figure 4. Development of (a) relative density, (b) εr value (measured in the ranges of 4.8~5.2 GHz), (c) Q × f value, and (d) τf value.
Materials 16 05195 g004
Figure 5. The bonding environment, coordinates, and charge distributions of cations in Sr2TiO4 ceramic.
Figure 5. The bonding environment, coordinates, and charge distributions of cations in Sr2TiO4 ceramic.
Materials 16 05195 g005
Figure 6. The bond ionicity f i μ value of chemical bonds in the Sr2TiO4 structure.
Figure 6. The bond ionicity f i μ value of chemical bonds in the Sr2TiO4 structure.
Materials 16 05195 g006
Figure 7. Contributions of chemical bonds to the lattice energy U value.
Figure 7. Contributions of chemical bonds to the lattice energy U value.
Materials 16 05195 g007
Table 1. Reported microwave dielectric material candidates with medium εr (20~70) values [5,6,7,8,9,10,11].
Table 1. Reported microwave dielectric material candidates with medium εr (20~70) values [5,6,7,8,9,10,11].
SystemST (°C)εrQ × f (GHz)τf (ppm/°C)Ref.
Nd1.02(Nb0.94Ta0.06)0.988O4127521.751,000−45.7[5]
Zn1.01Nb2O6/TiO2/Zn1.01Nb2O6120026.899,5000.5[6]
ZnTi0.7Ge0.3Nb2O8112035.662,700−58.0[7]
0.516ZTN–0.484ZNT110046.127,031−1.5[8]
Co(Ti0.8Zr0.2)Nb2O8125055.041,54159.8[9]
Sr3Ti2O7150063.084,000293.0[10]
Cu0.5Ti0.5NbO496071.211,00049.2[11]
Table 2. Atomic fractional coordinates of Sr2TiO4 ceramic system.
Table 2. Atomic fractional coordinates of Sr2TiO4 ceramic system.
AtomPositionFractional CoordinatesOcc 1Uiso 2
xyz
Sr4c0.0000.0000.355 (4)10.00988 (5)
Ti8d0.0000.0000.00010.00387 (6)
O18d0.0000.5000.00010.02071 (8)
O28d0.0000.0000.152 (3)10.02512 (4)
Bond lengthsSr–O1 × 4: 2.66932 (2) Å    Ti–O1 × 4: 1.94490 (10) Å
Sr–O2(1) × 4: 2.75192 (4) Å    Ti–O2 × 2: 1.91655 (3) Å
Sr–O2(2) × 1: 2.55960 (10) Å
1 Occ: site occupancy; 2 Uiso: isotropic atomic displacement parameters.
Table 3. Chemical bond susceptibility in the Sr2TiO4 structure.
Table 3. Chemical bond susceptibility in the Sr2TiO4 structure.
Bond Type N e μ E g μ   ( eV ) ZOμχμFμχμ/χ (%)
Sr–O1 × 40.193981 (7)5.66421 (4)47.415 (8) 0.333 27.12 (3)
Sr–O2(1) × 40.177033 (5)5.26030 (6)47.859 (3) 0.333 28.74 (8)
Sr–O2(2) × 10.220011 (8)6.26772 (3)46.854 (1) 0.083 6.27 (2)
Ti–O1 × 41.504495 (3)11.6008 (2)1213.933 (5) 0.167 25.48 (2)
Ti–O2 × 21.572252 (4)12.0245 (5)1213.547 (7) 0.083 12.39 (3)
Table 4. Thermal expansion coefficient in the Sr2TiO4 structure.
Table 4. Thermal expansion coefficient in the Sr2TiO4 structure.
Bond Type F m n μ U (kJ/mol) α m n μ ( 10 6 · K 1 ) α (10−6 K−1)
Sr–O1 × 40.333 2544 (8)20.23 (7)6.74 (1)
Sr–O2(1) × 40.333 2482 (3)20.81 (2)6.94 (3)
Sr–O2(2) × 10.083 658 (6)19.46 (5)1.62 (2)
Ti–O1 × 40.167 9849 (5)4.63 (4)0.77 (4)
Ti–O2 × 20.083 4978 (4)4.55 (1)0.38 (1)
αtotal (10−6 K−1) 16.45 (11)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, J.; Pang, J.; Luo, X.; Ao, L.; Xie, Q.; Wang, X.; Yang, H.; Tang, X. Phase Structure, Bond Features, and Microwave Dielectric Characteristics of Ruddlesden–Popper Type Sr2TiO4 Ceramics. Materials 2023, 16, 5195. https://doi.org/10.3390/ma16145195

AMA Style

Yang J, Pang J, Luo X, Ao L, Xie Q, Wang X, Yang H, Tang X. Phase Structure, Bond Features, and Microwave Dielectric Characteristics of Ruddlesden–Popper Type Sr2TiO4 Ceramics. Materials. 2023; 16(14):5195. https://doi.org/10.3390/ma16145195

Chicago/Turabian Style

Yang, Jun, Jinbiao Pang, Xiaofang Luo, Laiyuan Ao, Qiang Xie, Xing Wang, Hongyu Yang, and Xianzhong Tang. 2023. "Phase Structure, Bond Features, and Microwave Dielectric Characteristics of Ruddlesden–Popper Type Sr2TiO4 Ceramics" Materials 16, no. 14: 5195. https://doi.org/10.3390/ma16145195

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop